Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Article
  • Published:

Integrative analysis of ultra-deep RNA-seq reveals alternative promoter usage as a mechanism of activating oncogenic programmes during prostate cancer progression

Abstract

Transcription factor (TF) proteins regulate gene activity by binding to regulatory regions, most importantly at gene promoters. Many genes have alternative promoters (APs) bound by distinct TFs. The role of differential TF activity at APs during tumour development is poorly understood. Here we show, using deep RNA sequencing in 274 biopsies of benign prostate tissue, localized prostate tumours and metastatic castration-resistant prostate cancer, that AP usage increases as tumours progress and APs are responsible for a disproportionate amount of tumour transcriptional activity. Expression of the androgen receptor (AR), the key driver of prostate tumour activity, is correlated with elevated AP usage. We identified AR, FOXA1 and MYC as potential drivers of AP activation. DNA methylation is a likely mechanism for AP activation during tumour progression and lineage plasticity. Our data suggest that prostate tumours activate APs to magnify the transcriptional impact of tumour drivers, including AR and MYC.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: Activation and upregulation of APs are associated with increased expression of disease-related genes during PCa progression.
Fig. 2: FOXA1 binding and androgen signalling are associated with AP usage in PCa.
Fig. 3: MYC is a potential driver of AP activation in mCRPC.
Fig. 4: AP usage reflects lineage plasticity in response to therapy.
Fig. 5: Activation of APs is associated with DNA hypomethylation.

Similar content being viewed by others

Data availability

RNA-seq and whole-genome bisulfite sequencing data that support the findings of this study have been deposited in the European Genome–Phenome Archive (EGA) under accession code EGAS00001006275, and the SRA repository with Bioproject number PRJNA479544. Previously published RNA-seq data that were re-analysed here are available under accession code GSE115414, EGAD00001004424 and GSE119757. All other data supporting the findings of this study are available from the corresponding author on reasonable request.

Code availability

The R scripts and source data used to reproduce all figures and tables in this manuscript are available via GitHub at https://github.com/DavidQuigley/WCDT_alternative_promoter. The source data to reproduce all the figures and tables are available via Zenodo at https://zenodo.org/records/10966958.

References

  1. Carninci, P. et al. Genome-wide analysis of mammalian promoter architecture and evolution. Nat. Genet. 38, 626–635 (2006).

    Article  CAS  PubMed  Google Scholar 

  2. Landry, J. R., Mager, D. L. & Wilhelm, B. T. Complex controls: the role of alternative promoters in mammalian genomes. Trends Genet. 19, 640–648 (2003).

    Article  CAS  PubMed  Google Scholar 

  3. Demircioğlu, D. et al. A pan-cancer transcriptome analysis reveals pervasive regulation through alternative promoters. Cell 178, 1465–1477.e1417 (2019).

    Article  PubMed  Google Scholar 

  4. Davuluri, R. V., Suzuki, Y., Sugano, S., Plass, C. & Huang, T. H. The functional consequences of alternative promoter use in mammalian genomes. Trends Genet. 24, 167–177 (2008).

    Article  CAS  PubMed  Google Scholar 

  5. Greenberg, M. V. C. & Bourc’his, D. The diverse roles of DNA methylation in mammalian development and disease. Nat. Rev. Mol. Cell Biol. 20, 590–607 (2019).

    Article  CAS  PubMed  Google Scholar 

  6. Zhao, S. G. et al. The DNA methylation landscape of advanced prostate cancer. Nat. Genet. 52, 778–789 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Robinson, D. et al. Integrative clinical genomics of advanced prostate cancer. Cell 161, 1215–1228 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Quigley, D. A. et al. Genomic hallmarks and structural variation in metastatic prostate cancer. Cell 174, 758–769.e759 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Beltran, H. et al. Divergent clonal evolution of castration-resistant neuroendocrine prostate cancer. Nat. Med. 22, 298–305 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Nouruzi, S. et al. ASCL1 activates neuronal stem cell-like lineage programming through remodeling of the chromatin landscape in prostate cancer. Nat. Commun. 13, 2282 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Hua, J. T. et al. Risk SNP-mediated promoter-enhancer switching drives prostate cancer through lncRNA PCAT19. Cell 174, 564–575.e518 (2018).

    Article  CAS  PubMed  Google Scholar 

  12. Aggarwal, R. et al. Clinical and genomic characterization of treatment-emergent small-cell neuroendocrine prostate cancer: a multi-institutional prospective study. J. Clin. Oncol. 36, 2492–2503 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Aggarwal, R. R. et al. Whole-genome and transcriptional analysis of treatment-emergent small-cell neuroendocrine prostate cancer demonstrates intraclass heterogeneity. Mol. Cancer Res 17, 1235–1240 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Pinskaya, M. et al. Reference-free transcriptome exploration reveals novel RNAs for prostate cancer diagnosis. Life Sci. Alliance https://doi.org/10.26508/lsa.201900449 (2019).

  15. Chen, S. et al. Widespread and Functional RNA circularization in localized prostate cancer. Cell 176, 831–843.e822 (2019).

    Article  CAS  PubMed  Google Scholar 

  16. Pomerantz, M. M. et al. Prostate cancer reactivates developmental epigenomic programs during metastatic progression. Nat. Genet. 52, 790–799 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Yates, A. D. et al. Ensembl 2020. Nucleic Acids Res. 48, D682–D688 (2020).

    CAS  PubMed  Google Scholar 

  18. Jinesh, G. G. & Kamat, A. M. RalBP1 and p19-VHL play an oncogenic role, and p30-VHL plays a tumor suppressor role during the blebbishield emergency program. Cell Death Discov. 3, 17023 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Dai, C., Heemers, H. & Sharifi, N. Androgen signaling in prostate cancer. Cold Spring Harb. Perspect. Med. https://doi.org/10.1101/cshperspect.a030452 (2017).

  20. Jin, H. J., Zhao, J. C., Wu, L., Kim, J. & Yu, J. Cooperativity and equilibrium with FOXA1 define the androgen receptor transcriptional program. Nat. Commun. 5, 3972 (2014).

    Article  CAS  PubMed  Google Scholar 

  21. Puig, R. R., Boddie, P., Khan, A., Castro-Mondragon, J. A. & Mathelier, A. UniBind: maps of high-confidence direct TF-DNA interactions across nine species. BMC Genomics 22, 482 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Chen, W. S. et al. Genomic drivers of poor prognosis and enzalutamide resistance in metastatic castration-resistant prostate cancer. Eur. Urol. 76, 562–571 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Qiu, X. et al. MYC drives aggressive prostate cancer by disrupting transcriptional pause release at androgen receptor targets. Nat. Commun. 13, 2559 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Tran, M. G. B. et al. Independence of HIF1a and androgen signaling pathways in prostate cancer. BMC Cancer 20, 469 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Lin, C. Y. et al. Transcriptional amplification in tumor cells with elevated c-Myc. Cell 151, 56–67 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Barfeld, S. J. et al. c-Myc antagonises the transcriptional activity of the androgen receptor in prostate cancer affecting key gene networks. EBioMedicine 18, 83–93 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  27. Wang, J. et al. EZH2 noncanonically binds cMyc and p300 through a cryptic transactivation domain to mediate gene activation and promote oncogenesis. Nat. Cell Biol. 24, 384–399 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Kim, J. et al. Polycomb- and methylation-independent roles of EZH2 as a transcription activator. Cell Rep. 25, 2808–2820.e2804 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Serresi, M. et al. Functional antagonism of chromatin modulators regulates epithelial-mesenchymal transition. Sci. Adv. https://doi.org/10.1126/sciadv.abd7974 (2021).

  30. van Leenders, G. J. et al. Polycomb-group oncogenes EZH2, BMI1, and RING1 are overexpressed in prostate cancer with adverse pathologic and clinical features. Eur. Urol. 52, 455–463 (2007).

    Article  PubMed  Google Scholar 

  31. Guo, B. H. et al. Bmi-1 promotes invasion and metastasis, and its elevated expression is correlated with an advanced stage of breast cancer. Mol. Cancer 10, 10 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Lundberg, A. et al. The genomic and epigenomic landscape of double-negative metastatic prostate cancer. Cancer Res. https://doi.org/10.1158/0008-5472.Can-23-0593 (2023).

    Article  PubMed  PubMed Central  Google Scholar 

  33. Lei, J. & Howard, M. J. Targeted deletion of Hand2 in enteric neural precursor cells affects its functions in neurogenesis, neurotransmitter specification and gangliogenesis, causing functional aganglionosis. Development 138, 4789–4800 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Hendershot, T. J. et al. Conditional deletion of Hand2 reveals critical functions in neurogenesis and cell type-specific gene expression for development of neural crest-derived noradrenergic sympathetic ganglion neurons. Dev. Biol. 319, 179–191 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Mitchell, P. J., Timmons, P. M., Hébert, J. M., Rigby, P. W. & Tjian, R. Transcription factor AP-2 is expressed in neural crest cell lineages during mouse embryogenesis. Genes Dev. 5, 105–119 (1991).

    Article  CAS  PubMed  Google Scholar 

  36. Sjöström, M. et al. The 5-hydroxymethylcytosine landscape of prostate cancer. Cancer Res. https://doi.org/10.1158/0008-5472.can-22-1123 (2022).

  37. Shukla, S. K. et al. MUC1 and HIF-1alpha signaling crosstalk induces anabolic glucose metabolism to impart gemcitabine resistance to pancreatic cancer. Cancer Cell 32, 71–87.e77 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Lau, H. H., Ng, N. H. J., Loo, L. S. W., Jasmen, J. B. & Teo, A. K. K. The molecular functions of hepatocyte nuclear factors—in and beyond the liver. J. Hepatol. 68, 1033–1048 (2018).

    Article  CAS  PubMed  Google Scholar 

  39. Bonham, K., Ritchie, S. A., Dehm, S. M., Snyder, K. & Boyd, F. M. An alternative, human SRC promoter and its regulation by hepatic nuclear factor-1alpha. J. Biol. Chem. 275, 37604–37611 (2000).

    Article  CAS  PubMed  Google Scholar 

  40. Ci, X. et al. Heterochromatin protein 1α mediates development and aggressiveness of neuroendocrine prostate cancer. Cancer Res. 78, 2691–2704 (2018).

    Article  CAS  PubMed  Google Scholar 

  41. Qamra, A. et al. Epigenomic promoter alterations amplify gene isoform and immunogenic diversity in gastric adenocarcinoma. Cancer Discov. 7, 630–651 (2017).

    Article  CAS  PubMed  Google Scholar 

  42. Zhang, M., Moreno-Rodriguez, T. & Quigley, D. A. in Cancer Discovery Vol. 12, 2017−2019 (American Association for Cancer Research, 2022).

  43. Linder, S. et al. Drug-induced epigenomic plasticity reprograms circadian rhythm regulation to drive prostate cancer toward androgen independence. Cancer Discov. 12, 2074–2097 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Labbé, D. P. et al. TOP2A and EZH2 provide early detection of an aggressive prostate cancer subgroup. Clin. Cancer Res. 23, 7072–7083 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  45. Xu, K. et al. EZH2 oncogenic activity in castration-resistant prostate cancer cells is polycomb-independent. Science 338, 1465–1469 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Anwar, T., Gonzalez, M. E. & Kleer, C. G. Noncanonical functions of the polycomb group protein EZH2 in breast cancer. Am. J. Pathol. 191, 774–783 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Beltran, H. et al. The role of lineage plasticity in prostate cancer therapy resistance. Clin. Cancer Res. 25, 6916–6924 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Ferguson, A. M. & Rubin, M. A. Lineage plasticity in prostate cancer: looking beyond intrinsic alterations. Cancer Lett. 548, 215901 (2022).

    Article  CAS  PubMed  Google Scholar 

  49. Kulis, M., Queirós, A. C., Beekman, R. & Martín-Subero, J. I. Intragenic DNA methylation in transcriptional regulation, normal differentiation and cancer. Biochim. Biophys. Acta 1829, 1161–1174 (2013).

    Article  CAS  PubMed  Google Scholar 

  50. Kim, J. H. et al. Deep sequencing reveals distinct patterns of DNA methylation in prostate cancer. Genome Res. 21, 1028–1041 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Martin, M. Cutadapt removes adapter sequences from high-throughput sequencing reads. EMBnet.journal 17, 10–12 (2011).

    Article  Google Scholar 

  52. Dobin, A. et al. STAR: ultrafast universal RNA-seq aligner. Bioinformatics 29, 15–21 (2013).

    Article  CAS  PubMed  Google Scholar 

  53. Risso, D., Schwartz, K., Sherlock, G. & Dudoit, S. GC-content normalization for RNA-Seq data. BMC Bioinf. 12, 480 (2011).

    Article  CAS  Google Scholar 

  54. Heinz, S. et al. Simple combinations of lineage-determining transcription factors prime cis-regulatory elements required for macrophage and B cell identities. Mol. Cell 38, 576–589 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Frankish, A. et al. GENCODE reference annotation for the human and mouse genomes. Nucleic Acids Res. 47, D766–D773 (2019).

    Article  CAS  PubMed  Google Scholar 

  56. O’Leary, N. A. et al. Reference sequence (RefSeq) database at NCBI: current status, taxonomic expansion, and functional annotation. Nucleic Acids Res. 44, D733–D745 (2016).

    Article  PubMed  Google Scholar 

  57. Love, M. I., Huber, W. & Anders, S. Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome Biol. 15, 550 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  58. Lizio, M. et al. Gateways to the FANTOM5 promoter level mammalian expression atlas. Genome Biol. 16, 22 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Anders, S., Reyes, A. & Huber, W. Detecting differential usage of exons from RNA-seq data. Genome Res. 22, 2008–2017 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Liao, Y., Smyth, G. K. & Shi, W. featureCounts: an efficient general purpose program for assigning sequence reads to genomic features. Bioinformatics 30, 923–930 (2014).

    Article  CAS  PubMed  Google Scholar 

  61. Yu, G., Wang, L. G., Han, Y. & He, Q. Y. clusterProfiler: an R package for comparing biological themes among gene clusters. Omics 16, 284–287 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Liberzon, A. et al. The Molecular Signatures Database (MSigDB) hallmark gene set collection. Cell Syst. 1, 417–425 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Pico, A. R. et al. WikiPathways: pathway editing for the people. PLoS Biol. 6, e184 (2008).

    Article  PubMed  PubMed Central  Google Scholar 

  64. Tomlins, S. A. et al. Integrative molecular concept modeling of prostate cancer progression. Nat. Genet. 39, 41–51 (2007).

    Article  CAS  PubMed  Google Scholar 

  65. Sheffield, N. C. & Bock, C. LOLA: enrichment analysis for genomic region sets and regulatory elements in R and Bioconductor. Bioinformatics 32, 587–589 (2016).

    Article  CAS  PubMed  Google Scholar 

  66. Wu, H. et al. Detection of differentially methylated regions from whole-genome bisulfite sequencing data without replicates. Nucleic Acids Res. 43, e141 (2015).

    PubMed  PubMed Central  Google Scholar 

  67. Burger, L., Gaidatzis, D., Schubeler, D. & Stadler, M. B. Identification of active regulatory regions from DNA methylation data. Nucleic Acids Res. 41, e155 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  68. R Core Team. R: a language and environment for statistical computing. R Foundation for Statistical Computing https://www.R-project.org/ (2021).

Download references

Acknowledgements

We thank the patients who selflessly contributed samples to this study and without whom this research would not have been possible. This research was supported by a Stand Up To Cancer—Prostate Cancer Foundation Prostate Cancer Dream Team Award (SU2C-AACR-DT0812 to E.J.S.) and by the Movember Foundation. Stand Up To Cancer is a division of the Entertainment Industry Foundation. This research grant was administered by the American Association for Cancer Research, the scientific partner of SU2C. R.A. and M.S. were funded by a Prostate Cancer Foundation Young Investigator Award. D.A.Q. was funded by Young Investigator and Challenge awards from the PCF, by the UCSF Benioff Initiative for Prostate Cancer Research and by the United States Department of Defense (HT9425-23-PCRP-DSA). F.Y.F. was funded by Prostate Cancer Foundation Challenge Awards. Additional funding was provided by a UCSF Benioff Initiative for Prostate Cancer Research award. F.Y.F. was supported by National Institutes of Health (NIH)/National Cancer Institute (NCI) 1R01CA230516-01. F.Y.F. was supported by NIH/NCI 1R01CA227025 and Prostate Cancer Foundation (PCF) 17CHAL06. F.Y.F. was supported by NIH P50CA186786.

Author information

Authors and Affiliations

Authors

Consortia

Contributions

The studies were conceptualized and designed by M.Z., F.Y.F. and D.A.Q. Data analysis was carried out by M.Z., M.S., R.S., A.L. and H.X.D. Validation experiments were performed by X.C. and H.L. Biopsy samples were processed by A.F. and K.F. Resources were contributed by P.G.F., R.A., J.J.A., E.J.S. and the SU2C/PCF West Coast Prostate Cancer Dream Team. Supervision was provided by C.A.M., F.Y.F. and D.A.Q. The first draft of the manuscript was written by M.Z., M.S. and D.A.Q. All authors revised and approved the manuscript.

Corresponding author

Correspondence to David A. Quigley.

Ethics declarations

Competing interests

J.J.A. has consulted for or held advisory roles at Astellas Pharma, Bayer and Janssen Biotech Inc. He has received research funding from Aragon Pharmaceuticals Inc., Astellas Pharma, Novartis, Zenith Epigenetics Ltd. and Gilead Sciences Inc. F.Y.F. has consulted for Astellas, Bayer, Blue Earth Diagnostics, BMS, EMD Serono, Exact Sciences, Foundation Medicine, Janssen Oncology, Myovant, Roivant and Varian, and serves on the Scientific Advisory Board for BlueStar Genomics and SerImmune. F.Y.F. has patent applications with Decipher Biosciences on molecular signatures in PCa unrelated to this work. F.Y.F. has a patent application licensed to PFS Genomics/Exact Sciences. F.Y.F. has patent applications with Celgene unrelated to this work. The remaining authors declare no competing interests.

Peer review

Peer review information

Nature Cell Biology thanks Ramana Davuluri and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

Additional information

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

A full list of members and their affiliations appears in the Supplementary Information.

Extended data

Extended Data Fig. 1 Optimization for promoter activity estimation.

a) Overall schematic of the analysis. PAIR: from the Henri Mondor institution, CPCG: Canadian Prostate Cancer Genome Network, WCDT: West Coast Dream Team, t-SCNC: treatment-emergent small cell neuroendocrine carcinoma. b) An illustration of the promoter activity estimation methods. Solid boxes represent exons while the lines represent introns. The promoters (P1, P2, P3) are defined as the first 5’ TSSs (transcription start sites) of overlapping first exons. The splice junction reads (SJ) from the overlapping first exons were summed and log2-normalized to represent the promoter activities. The activity of the internal promoter P2 can be corrected by the split read ratios or split read subtractions method to exclude transcriptional activity from isoform B (TXB) (see Methods for details). TX: transcript, SJ: splice junction. c) Correlations between the CAGE (cap analysis of gene expression) tag reads and the promoter activities calculated using RNA-seq data of non-internal and internal promoters without and with corrections. (Spearman’s correlation test, two-sided) The matching CAGE and RNA-seq data from the same samples were from FANTOM5. Upper row: representative correlation plots showing one human adult testis sample. Lower row: a box plot showing Spearman’s correlation coefficients for all 67 samples with matching CAGE and RNA-seq data. Box plots show data from the 25th to the 75th percentile, with the median as a line inside the box. Whiskers extend to 1.5 times the interquartile range (IQR) from the lower and upper quartiles. d) Number of high confidence promoters (see Methods for details) in the non-internal and internal category. e) A representative sample down sampled to 31.25 million (M), 62.5 M, 125 M, 250 M, 500 M, and 750 M reads from 1000 M. The bars show the number of active promoters detected at each read depth (left y axis). Points show the number of new promoters detected per million reads (right y axis).

Extended Data Fig. 2 Activation of additional promoters is associated with gene expression upregulation.

a) The number of active promoters normalized to the number of expressed genes for each individual sample grouped by disease stages (N = 8, 147, 104 for benign, localized and mCRPC). Genes with nonzero counts were considered as expressed. Box plots show data from the 25th to the 75th percentile, with the median as a line inside the box. Whiskers extend to 1.5 times the interquartile range (IQR) from the lower and upper quartiles. b) Upregulated and downregulated genes were identified by differential gene expression analysis. Bar plot shows the percentage of genes in each category that switch between single-promoter active and multiple-promoter active in benign prostate and localized PCa (left) or mCRPC (right). Activated: switch from SP (single-promoter active) in benign to MP (multiple-promoter active) in tumors. Deactivated: switch from MP in benign to SP in tumors. Inactive: SP in both benign and tumors. Constitutively active: MP in both benign and tumors. c) The RNA-seq coverage across gene body from 5’ to 3’ for ten random samples from each of the dataset (PAIR, CPCG, and WCDT) in our data collection. d) The EDASeq bias plot of the positional biases in unnormalized promoter counts of all samples from the RNA-seq datasets (PAIR, CPCG, and WCDT) in our data collection. e) The analysis of number of genes switching from single promoter active in benign prostate to multiple promoters active in localized (left) or mCRPC (right) using the RNA-seq dataset all down-sampled to 80 M reads/sample. SP: single -promoter active, MP: multiple-promoter active. (Fisher’s exact tests, two-sided). f. Principal component analysis of all samples of different disease stages from three cohorts using the down-sampled RNA-seq dataset.

Extended Data Fig. 3 Alternative promoter usage occurs in cancer related genes.

a) Density plot of the Spearman’s correlation rho values between absolute promoter activity and corresponding gene expression for upregulated APs, downregulated APs and non-differential promoters in genes with differential APs in localized PCa vs benign prostate. b) Pathway enrichment analysis of genes with upregulated APs in mCRPC vs benign. Highlighted in red are pathways enriched for the genes with upregulated APs but not in upregulated genes. P values were calculated using hypergeometric tests. X axis shows the p values without multi-test adjustments, but the coloring was based on Benjamini Hochberg (BH)-adjusted p values. Dashed line shows unadjusted p value 0.05. c) Pathway enrichment analysis result of genes upregulated in mCRPC vs benign prostate. P values were calculated using hypergeometric tests. X axis shows the p values without multi-test adjustments, but the coloring was based on BH-adjusted p values. Dashed line shows unadjusted p value 0.05.

Extended Data Fig. 4 Alternative promoter usage is associated with AR levels.

a) Correlation between the number of upregulated APs in individual mCRPC samples with AR expression levels. 95% confidence interval for the predictions from a linear model is displayed. (Spearman’s correlation test, two-sided). b) The percentage of AR and FOXA1 co-binding in the FOXA1 bound upregulated APs in localized PCa and mCRPC (Fisher’s exact test, two-sided).

Extended Data Fig. 5 Alternative promoter usage is associated with driver transcription factors.

a, b) Unibind results showing significance of overlap between transcription factor (TF) ChIP-seq peaks and upregulated APs in localized PCa (A) or mCRPC (B). Each dot represents one ChIP-seq dataset. TFs were ranked by the most significant ChIP-seq dataset. P values were calculated using Fisher’s exact tests. Y axis shows the p values without multi-test adjustments, but the horizontal dashed line shows the corresponding BH-adjusted p value 0.05. c) Pathway enrichment analysis of genes with APs upregulated in mCRPC vs benign prostate and overlapping with MYC ChIP-seq peaks in LNCaP cells. P values were calculated using hypergeometric tests. X axis shows the p values without multi-test adjustments, but the coloring was based on BH-adjusted p values. Dashed line shows unadjusted p value 0.05.

Extended Data Fig. 6 Enriched pathways in genes whose promoters are bound by MYC, EZH2 or both.

Pathway enrichment analyses of genes with promoters overlapping with EZH2 LNCaP ChIP-seq peaks only (A), with MYC LNCaP ChIP-seq peaks only (B), and with both MYC and EZH2 LNCaP ChIP-seq peaks (C). P values were calculated using hypergeometric tests. X axis shows the p values without multi-test adjustments, but the coloring was based on BH-adjusted p values. Dashed line shows unadjusted p value 0.05.

Extended Data Fig. 7 Alternative promoter usage reflects lineage plasticity in mCRPC.

a) Unibind results showing significance of overlap between TF ChIP-seq peaks and downregulated APs in t-SCNC vs adenocarcinoma mCRPC. Each dot represents one ChIP-seq dataset. TFs were ranked by the most significant ChIP-seq dataset. P values were calculated using Fisher’s exact tests. Y axis shows the p values without multi-test adjustments, but the horizontal dashed line shows the corresponding BH-adjusted p value 0.05. b) Histogram showing the distribution of gastrointestinal (GI) scores across mCRPC samples. Dashed line splits the fourth quartile vs others.

Extended Data Fig. 8 DNA methylation at alternative promoters is anticorrelated with their activity.

a) Correlation between the promoter activity fold change and methylation differences at differentially active APs between mCRPC t-SCNC and adenocarcinoma mCRPC. 95% confidence interval for the predictions from a linear model is displayed. (Spearman’s correlation test, two-sided). b) Unibind results showing significance of overlap between TF ChIP-seq peaks and upregulated APs in mCRPC t-SCNC vs adenocarcinoma that overlapped with differentially hypomethylated regions in t-SCNC. Each dot represents one ChIP-seq dataset. TFs were ranked by the most significant ChIP-seq dataset. P values were calculated using Fisher’s exact tests. Y axis shows the p values without multi-test adjustments, but the horizontal dashed line shows the corresponding BH-adjusted p value 0.05.

Supplementary information

Supplementary Information

This file contains a list of Stand Up 2 Cancer (SU2C)/Prostate Cancer Foundation (PCF) West Coast Prostate Cancer Dream Team members and affiliations.

Reporting Summary

Supplementary Table 1

Supplementary Tables 1–8.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zhang, M., Sjöström, M., Cui, X. et al. Integrative analysis of ultra-deep RNA-seq reveals alternative promoter usage as a mechanism of activating oncogenic programmes during prostate cancer progression. Nat Cell Biol 26, 1176–1186 (2024). https://doi.org/10.1038/s41556-024-01438-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41556-024-01438-3

This article is cited by

Search

Quick links

Nature Briefing: Cancer

Sign up for the Nature Briefing: Cancer newsletter — what matters in cancer research, free to your inbox weekly.

Get what matters in cancer research, free to your inbox weekly. Sign up for Nature Briefing: Cancer